• No results found

Interpreting the seasonal environmental history recorded by

N/A
N/A
Protected

Academic year: 2022

Share "Interpreting the seasonal environmental history recorded by"

Copied!
50
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

Interpreting the seasonal environmental history recorded by

1

Arctic bivalves

2

Vihtakari Mikko1,2,3,, Ambrose William G. Jr.2,4, Renaud Paul E.2,5, Locke William L. V4,

3

Carroll Michael L.2, Berge Jørgen1,5, Clarke Leon J.6, Cottier Finlo7, Hop Haakon3

4

1 Department of Arctic and Marine Biology, UiT The Arctic University of

5

Norway, N-9037 Tromsø, Norway

6

2 Akvaplan-niva, Fram Centre, N-9296 Tromsø, Norway

7

3 Norwegian Polar Institute, Fram Centre, N-9296 Tromsø, Norway

8

4 Department of Biology, Bates College, Lewiston, Maine 04240, USA

9

5 University Centre in Svalbard, N-9171 Longyearbyen, Norway

10

6 School of Science and the Environment, Faculty of Science and Engineering,

11

Manchester Metropolitan University, Manchester, M1 5GD, UK

12

7 Scottish Association for Marine Science, Scottish Marine Institute, Oban, Argyll

13

PA37 1QA, UK

14

∗E-mail: mikko.vihtakari@gmail.com

15

Keywords:Serripes groenlandicus;Ciliatocardium ciliatum; Laser-Ablation Inductively-Coupled-

16

Plasma Mass-Spectrometry; paleoclimatology; paleoceanography; bivalve mollusk shells;in situ

17

analyses; shell mineralogy

18

Abstract

19

Understanding rapid climate change in the Arctic and its ecosystem implications requires more

20

information on the environment at temporal resolutions and time-periods not available from

21

the instrumental records. Such information can be acquired through geochemical proxy records,

22

but sub-annual records are rare in the literature. We analyzed shell material of bivalve mol-

23

lusks (Serripes groenlandicus andCiliatocardium ciliatum) that were placed on oceanographic

24

moorings for one year in two Arctic fjords to assess the potential use of shell elemental ratios

25

as environmental proxies. Li/Ca, Mg/Ca, Li/Mg, Mn/Ca, Sr/Ca, Mo/Ca and Ba/Ca were de-

26

termined using Laser-Ablation Inductively-Coupled-Plasma Mass-Spectrometry. The mooring

27

exposure, combined with previously derived sub-annual shell growth models, allowed us to re-

28

late the elemental ratio patterns to oceanographic data (temperature, salinity, and fluorescence)

29

collected by instruments attached to the moorings. Shell Ba/Ca profiles were characterized by

30

abrupt peaks occurring 11 to 81 days after the phytoplankton bloom, as indicated by the sea-

31

water fluorescence index. Li/Ca and Mg/Ca values exhibited a logarithmic relationship with

32

shell growth rate, indicated by marginal R2 of 0.43 and 0.30, respectively. These ratios were

33

also linearly related to temperature, with marginal R2 of 0.15 and 0.17, respectively. Mn/Ca

34

and Sr/Ca ratios exhibited variability among individuals and their temporal pattern was likely

35

controlled by several unidentified factors. Mo/Ca patterns within the shells did not demon-

36

strate meaningful correlations with any mooring instrument data. Our results reflect complex

37

relationships between elemental ratios, bivalve metabolism, methodological limitations, and syn-

38

chronized environmental processes suggesting that none of the studied elemental ratios can be

39

used as all-encompassing proxies of seawater temperature, salinity, paleoproductivity, or shell

40

(2)

growth rate. Despite this, Ba/Ca and Li/Ca can likely be used as sub-annual temporal anchors

41

in further studies, as the deposition of these elements likely occurred simultaneously within each

42

fjord.

43

1 Introduction

44

The annual sea ice cover over the Arctic Ocean has declined by approximately 20 % since the

45

industrial revolution [data from Figure 4.3a in 1] with an accelerating rate over the last decade [2].

46

Such a reduction in sea-ice cover, together with other anthropogenic perturbations, is expected

47

to cause dramatic changes in Arctic marine ecosystems [2, 3]. Understanding and anticipating

48

these rapid changes requires information about the past climate at sufficient temporal resolution

49

and over longer time-periods than that usually provided by instrumental records [4]. Such

50

knowledge can be acquired by interpretation of geochemical proxy records, which can represent

51

long time scales [4–6]. Whereas records of environmental changes at longer than decadal time-

52

scales may indicate correlative relationships between climatic and biological patterns, combining

53

environmental and biotic data at sub-annual scales can help identify the ecological mechanisms

54

through which climate regulates biotic processes. Unfortunately, there are few sub-seasonal

55

high-resolution records presented in the literature due to a paucity of available data.

56

Shells of many filter-feeding bivalve mollusks are promising geochemical proxy archives due

57

to: 1) largely sedentary nature of bivalves, meaning that individuals record temporal rather

58

than spatial variability in seawater conditions; 2) distribution of bivalves across a wide variety

59

of habitats and latitudes [7]; 3) representation of bivalve shells in the geological record [7–10]; 4)

60

longevity of bivalves allowing longer than decadal proxy records per individual [11–13]; and 5)

61

regular growth patterns in bivalve shells that can be used to develop growth chronologies [14–

62

17]. Two common circumpolar bivalve species, the Greenland cockle (Serripes groenlandicus

63

Mohr, 1786) and the hairy cockle (Ciliatocardium ciliatum Fabricius, 1780), have been used

64

as environmental and climatic indicators in the previous studies [18–22]. They are long lived

65

species forming an aragonitic shell [23–25] with prominent annual growth lines deposited during

66

a slow winter shell growth period that is regulated by food availability [17, 19, 26]. Their shell

67

growth is further affected by temperature and often correlates with large scale climatic drivers

68

over annual to decadal scales [18, 20–22].

69

In theory, the environmental information stored in bivalve shells can be used to hind-cast sea-

70

water conditions with a sub-annual resolution based on geochemical proxies, such as element-to-

71

calcium ratios, that are sampled along chronologically deposited shell material [27–29]. Several

72

elemental ratios, such as Li/Ca [30, 31], Mg/Ca [32, 33], and Sr/Ca [34], have been suggested as

73

proxies of seawater temperature in bivalve shells, but these ratios are often affected by metabolic

74

and kinetic processes, and thus may be used as temperature proxies only for specific cases when

75

shell growth rate and seawater temperature are strongly intercorrelated [35, 36]. Lithium to

76

magnesium ratio could potentially be used to tease apart the metabolic effects in Li/Ca and

77

Mg/Ca [37]. The ratios of barium, manganese, molybdenum, and lithium to calcium have been

78

suggested as proxies of pelagic productivity [31, 38–40]. Barium to calcium provides one of the

79

most consistent elemental ratio signals in bivalve shells: Ba/Ca profiles are characterized by a

80

flat background signal that is periodically interrupted by sharp peaks in a wide range of species

81

(3)

across various habitats and latitudes [24, 38, 39, 41–48]. In addition to potentially representing

82

variability in primary productivity, Ba/Ca may indicate ambient seawater concentrations [49].

83

In contrast, manganese is often associated with shell precipitation rate and may also be influ-

84

enced by seawater redox conditions, and therefore shows variable patterns depending on species

85

[50–54]. Molybdenum, on the other hand, may be incorporated through diet, making Mo/Ca a

86

potential proxy of paleoproductivity [40, 49].

87

Consequently, the development of elemental ratios in bivalve shells as environmental proxies

88

could be valuable, especially in the Arctic where instrumental records are short or interrupted

89

and climate change is rapid [55]. Elemental ratio proxies in bivalve shells are, however, compli-

90

cated by metabolism as calcium carbonate mineralization does not occur directly from seawater,

91

but takes place in a chemically controlled space; the extrapallial cavity [56–58]. Interpretation

92

of these geochemical proxies is further complicated by shell growth rate, which varies through

93

the year [17] and appears to influence some element ratios [36]. Consequently, understanding

94

the sub-annual growth patterns is a fundamental prerequisite for using any shell-based proxy

95

at sub-annual resolution. Very few studies, and none in the Arctic, have been able to relate

96

elemental ratios measured within bivalve shells to seawater parameters data recorded at the

97

growth location with sub-annual resolution.

98

In this study, we examine minor and trace elemental ratios within the shells ofS. groen-

99

landicusandC. ciliatum, and assess their potential use as environmental proxies. We deployed

100

these bivalves on moorings in two oceanographically contrasting fjords in Svalbard for one year

101

[17, 26]. The bivalve deployment combined with previously obtained sub-annual growth models

102

[17] allowed us to relate the elemental ratio patterns to the oceanographic data recorded by

103

mooring instrumentation. We aimed to examine whether: 1) Li/Ca, Ba/Ca, Mn/Ca or Mo/Ca

104

could be used as proxies of primary productivity as has been suggested by other studies, 2)

105

Li/Ca, Mg/Ca, Li/Mg or Sr/Ca could be used as proxies of temperature or shell growth rate,

106

and 3) any of the above mentioned elemental ratios were deposited simultaneously in different

107

individuals indicating that they could be used as sub-annual chronological markers in the studied

108

species.

109

2 Materials and Methods

110

2.1 Study design

111

A suite of element (Li, Mg, Mn, Sr, Mo, and Ba) to calcium ratios was determined for sub-annual

112

patterns in shells of two bivalve species (Serripes groenlandicus andCiliatocardium ciliatum)

113

deployed on oceanographic moorings for one year during the periods September 2007–2008 and

114

September 2009–2010 in two fjords on Svalbard: Kongsfjorden and Rijpfjorden. These two fjords

115

are oceanographically different. Kongsfjorden is an Atlantic water-influenced open fjord, whereas

116

Rijpfjorden is a fjord with a sill (depth 100-200 m) that is influenced mainly by Arctic water

117

masses [59–62]. Kongsfjorden was ice-free throughout the field deployment with the exception

118

of occasional drift ice, whereas Rijpfjorden was covered by sea ice for 8 months (January 21–

119

September 16) in 2007–2008 [63], and for 5 months (February 15–July 21) in 2009-2010 [17]. The

120

bivalve deployment on moorings is described in detail by Ambrose Jret al.[26] and Vihtakari

121

et al.[17]. In brief, bivalves were collected from the western Barents Sea in August 2007 and

122

(4)

from Svalbardbanken in August 2009. They were held in flow-through seawater tanks for 1–

123

4 weeks at the University Centre in Svalbard and incubated in seawater with 125 mg L1 of

124

calcein dye for 24 h immediately before they were placed in 7 mm mesh plastic cages (hereafter

125

baskets) on the oceanographic moorings. The calcein mark was used as an absolute time marker

126

of deployment and was identified in sectioned shells using fluorescent imaging [see 17]. During

127

2009-2010, the bivalves were deployed to two water depths, 15 m (basket A) and 25 m (basket

128

B), while in 2007-2008 they were deployed only to 25 m (Table 1). The bivalves were deployed

129

in September each year and recovered one year later.

130

Bivalves collected from the moorings were sacrificed and shells then were embedded in epoxy

131

resin [as described in 26]. Embedded shells were cut into thick sections along the maximum

132

growth axis, as described in Vihtakari et al. [17], and the thick sections were polished to a

133

thickness of 2.0±0.1 mm. These thick sections then were transferred to a clean room, where

134

they were rinsed and brushed in Milli-Q water, sonicated for 5 min and rinsed again. Finally,

135

the thick sections were left to dry overnight before they were analyzed using Laser-Ablation

136

Inductively-Coupled-Plasma Mass-Spectrometry (hereafter LA-ICP-MS). Eleven shells were fur-

137

ther analyzed forin situδ18O values using secondary ion mass spectrometry (SIMS) to determine

138

sub-annual growth models [see 17]. Measured element ratio patterns determined for nine shells

139

that demonstrated adequate growth models were compared to weekly averages of seawater tem-

140

perature, salinity and fluorescence index records obtained from mooring instruments located

141

adjacent to bivalve baskets (Table 2, see 17 for details).

142

2.2 Elemental ratio analyses

143

LA-ICP-MS [64] was conducted at the Plasma Mass Spectrometry Facility, Woods Hole Oceano-

144

graphic Institute (MA, US), using a Thermo-Finnigan Element2 HR-ICP-MS coupled to a New

145

Wave Laser UP 193 nm excimer laser ablation system. A sequence of holes was ablated along

146

the middle of the shell thick section from the outer margin to the calcein line [see 17] using

147

95 s dwell time, 10 Hz repetition rate and 90% output power. The analysis was conducted in

148

2009 for 2007-2008 deployment specimens and in 2011 for 2009-2010 deployment individuals.

149

Magnesium (25Mg), calcium (48Ca), manganese (55Mn), strontium (88Sr) and barium (138Ba)

150

were analyzed in both years. Molybdenum (98Mo) and lithium (7Li) were added to the analysis

151

for 2009-2010 samples. Due to the low concentration of Mo in the CaCO3matrix, 2009-2010

152

shells had a larger ablation crater size [¯x= 87.5±0.7µm (SE), n = 612] compared to 2007-2008

153

samples [¯x = 42.0± 0.3µm (SE), n = 311]. The distance between laser holes [¯x = 104.1±

154

14.3 (SD)µm ] was kept constant between sessions and samples, and therefore the number of

155

ablation holes varied between 17 and 64 per analyzed shell depending on the length of annual

156

growth increment.

157

The signal intensity (counts per second) of the analyzed elements was monitored in an

158

Element2 low resolution mode during the LA-ICP-MS analyses. The recording of element signal

159

intensity was started approximately 10 s after initiating the laser ablation to clean the shell

160

surface of debris and to ensure that the ablation plume material had reached the ICP-MS. An

161

estimated value for each element was generated by averaging 50 signal intensity measurements

162

during the peak of material flow. Nitric acid (5 % HNO3) was used as a blank, ensuring a

163

constant flow of the acid into the ICP-MS. Every tenth sample analyzed was a blank. The

164

(5)

moving average of blanks was calculated and subtracted from the data. Since the analyzed

165

shell matrix was predominantly aragonite [23, 25], 48Ca was used as an internal standard by

166

normalizing all other elements to Ca concentration [65]. Two standards, Japanese Certified

167

Reference Material or “JpnCRM” [66] and FEBS-1 [67], were run as every tenth and twentieth

168

sample, respectively. These standards were used to correct for instrument drift and to calibrate

169

elemental ratios to cover all isotopes. FEBS-1 was used for Mn/Ca and Li/Ca and JpnCRM

170

for the other elemental ratios. The reference materials did not have a certified value for Mo.

171

Therefore, Mo/Ca concentrations are given as percentage of Mo/Ca maximum for each shell

172

and comparison of absolute Mo/Ca values was not possible

173

2.3 Datasets and statistical analyses

174

The position of the LA-ICP-MS holes was related to sub-annual growth lines and a measurement

175

axis that was related to the historical location of the shell margin using ImageJ [68] and sclero

176

package [69] for R software [70], as described in Vihtakariet al.[17]. The method also allowed

177

a spatial estimation of averaging error [71, 72]. Resulting LA-ICP-MS sample distances are

178

therefore expressed as mm from deployment (i.e. the calcein mark) along the measurement axis,

179

together with minimum and maximum extents for each LA-ICP-MS hole (Figures S1–S6).

180

Growth models for nine shells (three from each basket: KB, RA and RB, Table 1), based on

181

estimated daily growth trajectories for SIMSδ18O centroids (Figure 9 in 17), allowed comparison

182

of elemental ratio data to mooring instrument data (temperature, fluorescence index and salinity)

183

and modeled growth rate. The estimated temporal extent sampled by each LA-ICP-MS hole was

184

used to calculate average growth rate, temperature, salinity, and fluorescence index values that

185

were used as predictor variables in consequent regression models. The averages were calculated

186

using daily values. The relationship between element ratios (response variable in all models)

187

and shell growth rate was logarithmic, and therefore growth rates were log-transformed before

188

analyses.

189

Linear mixed-effect regression models (LMMs) were used to examine the overall relationships

190

in the dataset by using samples as random effects, assuming a random intercept and a constant

191

slope (see Table S3 and Text S1 for definitions of the models). In order to examine the overall

192

variance of each elemental ratio explained by each predictor variable, LMMs were run separately

193

with each non-transformed predictor variable (Model 1; Table S3). Marginal and conditional

194

R2values for LMMs for these models were calculated using MuMIn package [73] for R [70] and

195

the method described by Johnson [74]. Marginal R2 values were used as a measure of overall

196

variance explained by each response variable and to examine whether the proxy relationship was

197

constant among samples. To examine the overall relative importance of each predictor variable

198

and the direction of the linear relationship, all predictor variables were combined as fixed effects

199

into a same LMM (Model 2; Table S3). Response variables were log-transformed, and predictor

200

variables centered to their means and scaled to their standard deviations before running Model

201

2. The fixed effects (effects of each predictor variable to an elemental ratio) then were scaled

202

to the maximum absolute value of 95% confidence intervals resulting to a measure of relative

203

effect for each fixed effect. Linear mixed-effect models were calculated using the nlme package

204

[75]. The variability in relationships between response and predictor variables among individual

205

samples was examined using linear regression models fitted for each sample, response variable

206

(6)

and predictor variable separately (Model 3; Table S3).

207

Coefficients of variation (CV) for minimum and maximum elemental ratios over the mooring

208

deployment were used to assess among individual consistency of elemental ratios using all ana-

209

lyzed shells over two deployment periods (n = 30, Table 1). Correlations between elemental ratios

210

and predictor variables for regression models were examined using principal component analy-

211

sis [76] calculated on correlation matrices averaged over samples using Fisher z-transformation

212

[77–79]. These correlation matrices are presented in Table S4.

213

3 Results

214

3.1 Oceanographic conditions in the fjords

215

Kongsfjorden experienced warmer temperatures in 2007-2008 than in 2009-2010 (Figure 1): The

216

autumn (September to December) temperatures in Kongsfjorden were on average 1.0C higher

217

in 2007 compared to 2009, the winter (January to April) temperatures 1.7C warmer, and the

218

spring/summer (May to September) temperatures 2.6C warmer in 2008 compared to 2010. In

219

contrast, temperature differences between years varied in Rijpfjorden: The autumn (September

220

to November) temperatures in Rijpfjorden were also on average 1.0C higher in 2007 compared

221

to 2009, the winter (December to May) temperatures were almost equal between deployment

222

years, but the summer temperatures were on average 2.4C lower in 2008 compared to 2010. In

223

Kongsfjorden, temperature began to increase in May in both years. In 2007-2008, temperature

224

remained above zero, while in the winter of 2009-2010, temperature was generally below zero.

225

Temperature was recorded at two depths (15 and 25 m) in 2009-2010. Temperature differences

226

between depths were generally small, except during the summer stratification period, when

227

temperature at 15 m was approximately 1C higher than at 25 m. Rijpfjorden experienced

228

temperatures close to -1.7C from January until July (6 months) in 2007-2008 and from Jan-

229

uary until June (5 months) in 2009-2010. Temperature rose abruptly in mid-July 2010, whereas

230

in 2008 it started increasing in mid-May, but did not exceed 0 C. In 2009-2010, tempera-

231

tures were similar at both measured depths until late August, when the surface layer cooled by

232

approximately 3C relative to the deeper (25 m) layer.

233

In both fjords, the fluorescence index (FLI) was close to zero prior to a dramatic increase

234

during the spring (Figure 1). The first fluorescence peak occurred later (mid-June to mid-July)

235

in Rijpfjorden than in Kongsfjorden (mid-May to beginning of June). Salinity was relatively

236

stable in Kongsfjorden, with a range between 33.3 and 35.0 (Figure 1). Rijpfjorden experienced

237

variable salinity regime, related to melt water from sea ice, from July to December. Salinity

238

varied more in 2009-2010 (34.6-30.6) than in 2007-2008 (34.3-31.7), and was most variable at

239

the shallow baskets (15 m).

240

3.2 Patterns in element ratio profiles

241

Lithium to calcium ratios were consistently lower during winter and increased after the winter

242

growth band in all studied shells (Figures 2, S3–S6). The increase occurred simultaneously

243

with increased growth rate in growth modeled shells (Figures 2 and S7). Minimum Li/Ca was

244

13.9±0.3 (SE, n = 22)µmol mol−1on average (Table 3). The Li/Ca minimum was deposited

245

(7)

sometime between October and late May in Kongsfjorden and between October and mid-July

246

in Rijpfjorden (Figure 2). Coefficient of variation for minimum Li/Ca values varied between

247

7.5 and 14.1 % among baskets and was higher than that for maximum values (Table 3). The

248

maximum values were 21.6±0.3 (SE, n = 22) on average, and were estimated to occur July to

249

early September in Kongsfjorden and mid-July to early August in Rijpfjorden (Figure 2).

250

Magnesium to calcium ratios were at their lowest during the winter growth band and in-

251

creased immediately after or towards the end of the winter growth period in most analyzed

252

shells (Figures 2 and S1–S6). Three shells deployed to Rijpfjorden in 2007, however, did not

253

demonstrate clear seasonal Mg/Ca fluctuations (Figure S2). The strongest increase in Mg/Ca

254

values occurred during spring together with increased growth rate (Figures 2 and S8). After

255

reaching the maximum in July to mid-August in Kongsfjorden and in late July to late August in

256

Rijpfjorden, Mg/Ca values decreased slightly until the end of the deployment period (Figure 2).

257

Maximum Mg/Ca values ranged between 1.04 and 4.15 mmol mol1 being generally higher in

258

2009-2010 than in 2007-2008 (Table 3). Minimum Mg/Ca values ranged between 0.39 and 1.70

259

mmol mol1and were not obviously different among years. Coefficient of variation for Mg/Ca

260

minimum and maximum values was higher than that for Li/Ca (Table 3).

261

Manganese to calcium values exhibited variable patterns, but were also characterized by

262

peaks deposited during the translucent summer growth period in 24 of 30 analyzed shells (Figure

263

S1–S6). These peaks were deposited sometime between late May and August in Kongsfjorden,

264

and between early July and early August in Rijpfjorden occurring one to 70 days after the

265

fluorescence peak (Table 4 and Figure 2). Low Mn/Ca values were deposited during the winter

266

growth band from January until the end of the winter growth band (Figure S9). Average

267

maximum manganese values ranged between 1.31 and 8.52µmol mol1 (Table 3). Maximum

268

Mn/Ca values within baskets showed high variability as illustrated by coefficient of variation

269

(Table 3). Average minimum Mn/Ca values ranged between 0.16 and 0.75µmol mol1among

270

baskets, and coefficient of variation was high (Table 3). Average minimum and maximum values

271

were lower in 2009-2010 (Table 3).

272

Individuals within baskets demonstrated considerable variability with respect to Sr/Ca pro-

273

files (Figures 2, S1–S6). Minimum values were deposited before the winter growth band in 3

274

samples, during the winter growth in 4 samples, and after the winter growth in 23 samples.

275

Furthermore, maximum Sr/Ca values occurred before, during and after the winter growth band

276

in 7, 7, and 16 samples, respectively (Figures S1–S6). Minimum Sr/Ca values were deposited

277

between May and August in two growth modeledS. groenlandicus from Kongsfjorden and be-

278

tween October and March in the growth modeledC. ciliatumspecimen (Figures 2 and S10). In

279

Rijpfjorden, the minimum values were deposited between July and mid-August in seven shells

280

and between April and mid-July in oneS. groenlandicusspecimen (Figure 2). Maximum Sr/Ca

281

values in growth modeled shells from Kongsfjorden were deposited at the end of the mooring de-

282

ployment in mid-September, whereas Rijpfjorden shells showed more variability with maximum

283

values occurring in the beginning of the mooring deployment (September to December) as well

284

as towards the end of the mooring deployment (August to September, Figure 2). Coefficient

285

of variation for minimum and maximum Sr/Ca values was lower than those for Mg/Ca (Table

286

3). Minimum Sr/Ca value was 1.32±0.04 mmol mol−1(SE, n = 30) on average and maximum

287

value 2.37±0.09 mmol mol−1(SE, n = 30).

288

Molybdenum to calcium ratios were at their highest during or before the winter growth

289

(8)

band in all shells analyzed for Mo/Ca (2009-2010) and the ratios decreased after the end of

290

the growth check (Figures 2, S3–S6). After the minimum Mo/Ca, which occurred between

291

mid-April and September in Kongsfjorden and between July and August in Rijpfjorden, Mo/Ca

292

values increased again until the end of the mooring exposure (mid-September 2010, Figures 2

293

and S11). Maximum Mo/Ca values were measured at the beginning of the mooring deployment

294

(September to April, Figure 2).

295

Barium to calcium profiles were characterized by abrupt unimodal peaks (maximum values

296

= 3.1-76.1µmol mol1, ¯x= 20.0µmol mol1, Table 5) that were differentiated from low Ba/Ca

297

background levels (0.43–2µmol mol1, ¯x= 1µmol mol1, Figures 2, S1–S6). The peaks appeared

298

annually, occurring subsequent to the winter growth band in 27 of 30 analyzed shells (Figures

299

S1–S6). A distinct barium peak was not present in twoC. ciliatum from 2007-2008 deployed

300

in the 25 m basket in Rijpfjorden (Figure S2) and oneC. ciliatum from 2009-2010 deployed

301

in the 15 m basket in Kongsfjorden (Figure S3). In 2009-2010 samples, the Ba maxima were

302

considerably lower in the 25 m basket in Rijpfjorden compared to other baskets (RB in Table 3

303

and Figure S6). Barium peak values were not consistent within a basket as indicated by high

304

coefficient of variation (Table 3). The minimum Ba/Ca values were associated with a lower

305

within basket variability than the maximum values (Table 3). Barium peaks in Kongsfjorden

306

were estimated to occur between June and mid-August, 18 to 100 days after the fluorescence

307

peak (Table 5). Further, Ba/Ca peak values were deposited in July in Rijpfjorden occurring 11

308

to 36 days after the first peak in fluorescence index (Table 5).

309

3.3 Correlations between element ratios, growth rates and mooring instru-

310

ment data

311

Li/Ca and Mg/Ca covaried within 2009-2010 shells as indicated by arrows pointing approxi-

312

mately to the same direction in the PCA plot (Figure 3B) and high correlation coefficients (rz 313

= 0.78, r = 0.13–0.92; Table S4). Similar correlations between element ratios were evident for

314

Sr/Ca and Mo/Ca in 2009-2010 (Figure 3B,rz= 0.59, r =−0.69–0.99), Mn/Ca and Ba/Ca –

315

especially in the growth modeled shells (Figure 3C,rz= 0.50, r =−0.02–0.78), and Mg/Ca and

316

Mn/Ca in 2007-2008 shells (Figure 3A,rz= 0.38, r =−0.31–0.78). Further, Mg/Ca and Li/Mg

317

were strongly negatively correlated in 2009-2010 shells as demonstrated by arrows pointing to

318

opposite directions in the PCA plot (Figure 3B,rz=−0.92, r =−0.99 –−0.60). Also Li/Ca

319

and Li/Mg, Mg/Ca and Mo/Ca, and Li/Ca and Mo/Ca were negatively correlated (Table S4).

320

Temperature and salinity were negatively correlated (rz=−0.71, r =−0.86–0.57), whereas tem-

321

perature yielded positive correlations with fluorescence (rz= 0.48, r = 0.34–0.67) and logarithm

322

of shell growth rate (rz= 0.43, r = 0.22–0.66, Figure 3D).

323

Overall, logarithm of growth rate was the best explanatory factor for element ratio variability

324

in growth modeled shells (Figure 4A). Coefficient of determination (R2) for individual samples

325

ranged between 0.19 and 0.75 for the regression between Li/Ca and growth rate, between 0.30

326

and 0.59 for Mg/Ca, between 0.11 and 0.24 for Li/MG, and between 0.01 and 0.87 for Mn/Ca

327

(Table S2). Also Sr/Ca exhibited significant regressions with growth rate, but these relation-

328

ships varied from positive to negative (Table S2). Temperature yielded significant regressions

329

with Li/Ca, Mg/Ca, Li/Mg, and Sr/Ca (Table S2), but in the majority of samples these regres-

330

sions were not as strong as those for logarithm of growth rate (Figure 4A). The temperature

331

(9)

relationships for Li/Ca, Mg/Ca and Li/Mg were relatively consistent among samples, although

332

associated with large residual standard error (Tables S1–S2).

333

4 Discussion

334

Barium, manganese, molybdenum, and lithium to calcium ratios have previously been related to

335

primary production [31, 38–40] (Section 4.1). Although Mn/Ca and Ba/Ca exhibited patterns

336

that resembled the patterns in the fluorescence index (Figure 2), which was used as a proxy

337

of primary production, the differences in peak heights among samples from the same basket

338

suggested that these element ratios were also affected by other processes and could not be used

339

as straightforward proxies of primary production (Tables 3–5; see Section 4.1). Despite this,

340

Ba peaks were deposited likely at the same in a basket, but the timing varied between baskets

341

occurring 11 to 81 days after the phytoplankton bloom (Figure 2, Table 5, Section 4.3). Ba/Ca

342

could potentially be related to dissolved or particular Ba in ambient seawater. Mo/Ca and

343

Li/Ca did not exhibit patterns that could have been linked to primary production (Figure 2).

344

Lithium, magnesium and strontium to calcium ratios, in turn, have been suggested as proxies

345

of growth rate or temperature [30–34] (Section 4.2). We did observe considerable similarities

346

between Li/Ca, Mg/Ca, growth rate and temperature (Figures 3–4), but individual samples

347

from a same basket demonstrated variability in element-to-calcium ratios making it difficult to

348

use these ratios as proxies of absolute growth rate or temperature (see Section 4.2). Neverthe-

349

less, Li/Ca might reflect crystal growth rate in bivalve shells, whereas Mg/Ca appears to be

350

loosely linked with temperature (Figures 5–6). Finally, individuals within baskets demonstrated

351

variability in Sr/Ca profiles that could not satisfactorily be explained by any single predictor

352

variable (growth rate, temperature, fluorescence and salinity) used in this study (Figure 4).

353

In general, our results highlight the limitations caused by metabolically controlled deposition

354

of CaCO3in bivalves [56, 57] suggesting that none of the studied element ratio could be used

355

as straightforward proxies of temperature, salinity, paleoproductivity or shell growth rate. In

356

following sections we discuss the studied element ratios as potential proxies of primary produc-

357

tion (Section 4.1), shell growth rate or temperature (Section 4.2), and sub-seasonal temporal

358

anchors (Section 4.3). We also highlight the methodological constraints associated with our data

359

(Section 4.4).

360

4.1 Potential proxies of primary production

361

Barium to calcium profiles were characterized by distinct unimodal peaks, which resembled the

362

peaks in fluorescence index (Figures 1, 2, and S1-S6). The barium peak in Kongsfjorden shells

363

occurred approximately 74 days after the peak in phytoplankton bloom, which took place in

364

mid-May, and 19 days after ice-algae/phytoplankton associated fluorescence peak in Rijpfjor-

365

den (Table 5). Dissolved barium from seawater, which in turn is sometimes connected with

366

phytoplankton blooms [46, 80], has been found to consistently incorporate into calciticMytilus

367

edulisandPecten maximusshells with a partition coefficient of approximately 0.1 [39, 49]. Ap-

368

plied to our shells, Ba/Ca values should have been approximately similar, within the averaging

369

error framework (see Section 4.4), in each basket assuming that calcium was uniformly dis-

370

tributed along studied shells. Measured Ba/Ca background values varied between 0.4 and∼2

371

(10)

µmol mol1, were consistent with those reported earlier [46], and did not show any obvious vari-

372

ation within baskets that could not have been explained by averaging error (Table 3). Measured

373

maximum Ba/Ca values, on the other hand, varied between 3.1–76.1µmol mol1demonstrating

374

different peak values among shells from a same basket (Table 3). This variability in maximum

375

values is among the largest reported [46], and cannot completely be explained by averaging error

376

(see Section 4.4).

377

Predictor variables did not satisfactorily explain the Ba/Ca peaks: although Ba/Ca peaks

378

occurred simultaneously with increased shell growth in all growth modeled shells (Figures 2

379

and S12), growth rate explained only 2% of Ba/Ca variation across samples (marginal R2from

380

LMM; Figure 4) and<1 to 18% among samples (R2from regression models; Table S2). Further,

381

temperature was negatively related with Ba/Ca explaining 2% of variation across samples (Fig-

382

ure 4). Bivalve age, shell height, or length of the growth increment during mooring deployment

383

did not yield significant slopes in a regression model with Ba/Ca peak values, but Ba/Ca peak

384

values were significantly lower in the 25 m basket in Rijpfjorden compared to other baskets.

385

Therefore, our results are inconclusive about the environmental factors associated with the ob-

386

served barium peaks. Nevertheless, the considerable differences in Ba/Ca maximums among

387

samples from a same basket and the variable time-lag from bloom between fjords (Table 5) sug-

388

gest that although Ba/Ca might be connected to processes related to primary production, the

389

ratio cannot be used as a direct paleoproductivity proxy, agreeing with what has been suggested

390

by recent studies [45, 46, 49, 81].

391

In addition to barium, manganese to calcium profiles also demonstrated peaks that resembled

392

the fluorescence index peaks (Figures 1, 2 and S1-S6). Despite the seemingly synchronous

393

deposition of Mn in growth modeled shells (Table 4), Mn/Ca patterns exhibited individual

394

differences among shells from a same basket (Table 3). Further, Mn/Ca values in the growth

395

modeled shells were clearly correlated with growth rate (Figure 4 and Table S2) demonstrating

396

that Mn/Ca incorporation is likely, at least partly, kinetically controlled. Manganese occurs

397

partly as non-lattice-bound element in an aragonitic bivalveCorbula amurensis[82]. A varying

398

amount of Mn not directly bound to CaCO3matrix could also explain the mixed Mn/Ca results

399

in our study. Nevertheless, Mn/Ca peaks occurring approximately simultaneously in growth

400

modeled shells also demonstrate a degree of synchronous environmental or physiological control.

401

Previous studies suggest that Mn/Ca could partly be incorporated in relationship with Mn

402

concentration in seawater [51, 83]. Phytoplankton blooms have also been suggested as a cause

403

for Mn fluctuations in bivalve shells [24, 38]. Our data do not support the direct connection with

404

phytoplankton bloom events, but it is possible that pelagic Mn cycle is connected to productivity

405

to some extent as reviewed by [83]. Consequently, Mn/Ca is a potential, but complicated proxy

406

of several environmental and physiological factors in both species.

407

Maximum molybdenum to calcium values were measured during autumn before the depo-

408

sition of the winter growth band in all growth modeled shells (Figure 2). Consequently, our

409

dataset did not demonstrate prominent Mo peaks occurring during spring as has been reported

410

for calcitic scallopsComptopallium radula [48] andP. maximus[40]. Nevertheless, Mo/Ca pro-

411

files were relatively similar among shells demonstrating that Mo/Ca values either fell under the

412

detection limit of ICP-MS or that the incorporation mechanism could have been environmentally

413

regulated. The incorporation of Mo into bivalve shells might occur through diet, which makes

414

Mo/Ca a promising environmental proxy [40, 49]. If this was the case local phytoplankton may

415

(11)

not have been enriched in Mo. Alternatively, Mo could be connected to sediment surface redox-

416

processes [28] or sediment particles, as bivalves in our study were deployed in the water column

417

and did not grow in their natural habitat. Although our results do not preclude the possibility

418

for Mo/Ca being a potential proxy inS. groenlandicusandC. ciliatum, more research is needed

419

to draw further conclusions about this elemental ratio.

420

Our data did not demonstrate a clear connection between fluorescence index and Li/Ca

421

(Figures 4, and S7) casting a doubt on the hypothesis of phytoplankton blooms causing Li/Ca

422

peaks [31]. Therefore, Li/Ca peaks cannot be used as a proxy of timing and magnitude of

423

phytoplankton blooms in studied shells, although it is possible that phytoplankton blooms could

424

have contributed to increasing the Li/Ca values in Kongsfjorden (Table S2).

425

4.2 Potential proxies of growth rate or temperature

426

Lithium to calcium patterns were similar among individuals in baskets suggesting synchronized

427

responses to environmental or physiological processes (Figure 2 and Table 3). Logarithm of

428

average growth rate explained 43% of overall Li/Ca variation across all samples (LMM, Figure 4),

429

and 19–75% among samples (regressions, Table S2). Li/Ca–shell growth rate relationships were

430

logarithmic unlike in previous published studies where the authors reported linear relationships

431

with a similar slope forP. maximus[31] andArctica islandica[30] (Figure 5A). Shell growth rate

432

is an indicator of crystal growth rate in bivalve mollusk shells [31, 84]. Therefore, the positive

433

correlations between Li/Ca and shell growth rate agree with other published studies suggesting

434

that crystal growth rate is likely the primary driver of Li/Ca incorporation in bivalve mollusk

435

shells [30, 31]. Nevertheless, studies report differing regression equations between Li/Ca and

436

shell growth rate and these relationships do not yield particularly high R2values (Figure 5A).

437

This suggests that also other factors affect Li/Ca incorporation.

438

Temperature and riverine output have also been suggested to partly control Li/Ca in bivalve

439

shells [30, 31]. Since temperature and growth rate were correlated in our shells [17], the effects of

440

these factors are difficult to separate. Nevertheless, temperature significantly explained Li/Ca

441

variability, although these correlations were generally not as strong as for shell growth rate (Fig-

442

ures 4–5 and Table S2). The imprecision in our growth models could have contributed to the

443

lower temperature correlations, as a one-month shift in Li/Ca peak would have led to consider-

444

ably stronger temperature correlations for Rijpfjorden shells (Figures 2 and S7). Despite this,

445

the relationships for species that have been studied so far do not appear to demonstrate strong

446

enough R2values to reconstruct seawater temperatures (Figure 5B). Instead, significant regres-

447

sions between Li/Ca and temperature in bivalve mollusk shells (Figure 5B) could be explained

448

by dependency between temperature and shell growth rate, and therefore CaCO3crystal growth

449

rate.

450

Since we lack element concentration measurements in seawater, we can only speculate about

451

the effect of riverine output increasing Li concentration in ambient water and therefore contribut-

452

ing to shell Li/Ca [30]. Li/Ca peaks were coincident with decreased salinity (Figures 2 and S7).

453

If melt-water events increased Li concentration in ambient water in our study, it is possible that

454

these events could have contributed to Li/Ca fluctuations as suggested by Th´ebaultet al.[30].

455

Despite the uncertainties in our dataset, we can conclude, with a relatively high certainty, that

456

Li/Ca cannot be used as a temperature proxy inS. groenlandicus andC. ciliatumshells, but

457

(12)

appears to be a promising proxy of shell and/or crystal growth rate. Li/Ca, however, did not

458

yield strong enough relationships to precisely reconstruct sub-annual shell growth.

459

Relatively consistent patterns in Mg/Ca among individuals from the same basket (Figures 2,

460

S1–S6) suggested that the incorporation of Mg/Ca is likely related to synchronized environmental

461

or physiological processes. A large coefficient of variation, however, indicates that these processes

462

do not yield similar Mg/Ca peak values among shells (Table 3). Relatively strong correlations

463

with logarithm of average growth rate indicated that incorporation of Mg/Ca could be related to

464

shell precipitation rate similarly to Li/Ca (Figure 4). Furthermore, Mg/Ca correlated positively

465

with temperature (Figure 4 and Table S2). Many studies have reported similar significant

466

correlations between Mg/Ca ratio and sea surface temperature [32–35, 44, 53, 85–88]. Most of

467

these studies report either a large variability in temperature correlations similar to our study

468

[e.g. 34, 35], or that the relationship is restricted to certain conditions [e.g. 86, 87]. Organic

469

matter prior the elemental analysis has been removed in some studies that have reported strong

470

relationships between temperature and Mg/Ca [32, 89].

471

Our Mg/Ca–temperature relationships are similar to those reported for calcitic bivalves

472

Mytilus trossulus[32],M. edulis[90], andP. maximus[88] with the exception that coefficients of

473

variation are clearly lower in our study (Figure 6). Mg/Ca is thought to be strongly metabolically

474

controlled in marine bivalves: present day Mg/Ca molar ratio is 5.2 mol mol1[91], but report

475

Mg/Ca ratios in bivalve CaCO3 that are several orders of magnitude lower than the ambient

476

molar ratios (varied between 0.0041 and 0.0004 mol mol1in this study). Furthermore, Mg/Ca

477

is precipitated to inorganic aragonite following an inverse relationship with expected molar

478

ratio of>0.085 mol mol1 for the temperatures in this study [92]. Despite this, most reported

479

Mg/Ca–temperature relationships are positive (Figure 6),Crassostrea gigasbeing an exception

480

[54]. It should also be noted that Mg/Ca–temperature relationships appear generally stronger

481

for calcitic bivalves (bivalves in Figure 6) than for aragonitic bivalves (such asS. groenlandicus,

482

C. ciliatumandA. islandica[e.g. 35]). It seems feasible that Mg/Ca functions as a temperature

483

proxy in many bivalve shells (Figure 6), but Mg/Ca incorporation is also influenced by other

484

factors such that the imprecision associated with temperature estimates derived from Mg/Ca

485

is often larger than the seasonal temperature fluctuations. Our results are consistent with this

486

hypothesis and indicate that Mg/Ca is an unreliable temperature proxy forS. groenlandicus

487

andC. ciliatum. Nevertheless, our results also indicate that temperature does correlate with

488

Mg incorporation, and further studies should consider removal of organic matter before ICP-MS

489

analyses.

490

Studies on corals have demonstrated that combining Li/Ca and Mg/Ca could potentially

491

be used to tease apart the metabolic effects associated with these ratios and strengthen the

492

temperature relationship [37]. Our results, however, demonstrated generally weaker correlations

493

between Li/Mg and temperature than those between Li/Ca and temperature and Mg/Ca and

494

temperature separately (Figure 4, Table S2). Consequently, Li/Mg does not provide a robust

495

temperature proxy.

496

Strontium-to-calcium ratio was significantly affected by all predictor variables (Figure 4),

497

temperature and fluorescence index yielding the most consistent regressions (Table S2). Coeffi-

498

cient of variation for Sr/Ca maximum values indicates that Sr/Ca values varied among samples

499

from a same basket (Table 3). The large variability in Sr/Ca among samples from a same

500

location is consistent with the literature [44, 93] and suggests that any environmental signals

501

(13)

in Sr/Ca may be difficult to separate from vital effects. Strontium partition into calcium car-

502

bonate is related to the crystal growth rate of CaCO3matrix [92, 94]. Although, some earlier

503

studies have successfully used Sr/Ca as a temperature proxy [85, 95, 96], more recent studies

504

question the relationship [50, 97, 98]: it seems possible that temperature and crystal growth

505

rate of CaCO3 skeleton are connected resulting in a positive correlation between Sr/Ca and

506

temperature. Judging from our data, this was not the case for studied shells.

507

4.3 Sub-seasonal temporal anchors

508

Barium-to-calcium maximum values were deposited at approximately same time among samples

509

from the same basket (Table 5) considering the uncertainty caused by LA-ICP-MS averaging

510

error and growth models derived fromδ18O values (see Section 4.4). Measured Ba/Ca maximums

511

were estimated to be deposited in mid-July to early August in Kongsfjorden (Table 5). Barium

512

peaks in Rijpfjorden occurred during or right after a fast shell growth period (Figures 2 and S12)

513

and were timed to occur early July in the basket at 15 m depth and late July, 12 days later, in

514

the deeper basket at 25 m depth (Table 5). Simultaneous occurrence of Ba/Ca maximums within

515

baskets and similar patterns in 29 of 32 analyzed shells (Figure S1-S6) indicates synchronous

516

environmental or physiological drivers for incorporation of Ba in studied shells. Synchronously

517

deposited chemical proxies are useful temporal anchors to combine chronologies across bivalves

518

sampled from the same location [29]. Our results indicate that the Ba/Ca peaks are likely to

519

occur simultaneously 2.5 months to 2.5 weeks after primary production bloom, and they can be

520

used as sub-annual anchors across shells from a same location, if averaging error of elemental

521

sampling is kept sufficiently low.

522

Li/Ca also demonstrated remarkably synchronous patterns within baskets (Table 3) as min-

523

imum and maximum value variability could likely be explained by averaging error caused by

524

LA-ICP-MS sampling (see Section 4.4). Therefore, Li/Ca peak and trough values could have

525

been approximately similar across individuals from a same basket further demonstrating the

526

synchronized incorporation of this element ratio. Overall, Li/Ca ratios corresponded with those

527

reported by Th´ebaultet al.[30]: the range of Li/Ca fluctuation they reported was 1.3 to 1.6

528

fold over a growing season, whereas lithium values in this study varied between 1.3 and 2.2 fold

529

(1.6 on average). This demonstrates that Li/Ca could work as a temporal anchor also for other

530

species thanS. groenlandicus andC. ciliatum. Since Li/Ca peaks were rather broad in studied

531

shells it is advisable to use the increases in Li/Ca as temporal anchors.

532

4.4 Methodological limitations

533

The bivalves in this study were held in the water column on oceanographic moorings, and

534

therefore they might not have recorded elemental ratios similarly to their natural habitat. The

535

mooring deployment likely excluded the effect of sediment-surface redox-processes, which have

536

been suggested as important contributors for the seasonal dynamics of, at least, Mn [28, 48,

537

83, 99]. Further, we did not observe similar seasonal patterns in Sr/Ca ratios that has been

538

reported earlier forS. groenlandicus[18, 24]. It is possible that Sr/Ca is partly connected with

539

sediment surface processes and therefore our shells did not record all possible variability for this

540

element ratio.

541

(14)

The extent of time averaging sampled by LA-ICP-MS is relative to the sample volume and

542

average shell growth rate over the sampled area [71, 72]. Because sample hole size in our study

543

varied little within years (see Section 2.2), time averaging was related to shell growth rate. Even

544

though LA-ICP-MS sampling was able to capture the Ba/Ca peaks (Figures 2, S1–S6) it is

545

possible that time-averaging contributed to profiles of some elements during low growth rate

546

such that no meaningful environmental correlations were found [100].

547

Growth models used to determine the time extent for each LA-ICP-MS sample were subject

548

to uncertainty [17]. It is unlikely that these growth models were an entirely accurate representa-

549

tion of the actual growth during the mooring deployment and, therefore, our dataset contained a

550

bias, which increased correlations between element ratios and average shell growth rate, because

551

shell growth rate was obtained from growth models, which affected the alignment of elemental

552

ratios. Further, shell growth rate and temperature were significantly correlated in all growth

553

modeled shells (Figure 3; 17).

554

Even though we attempted to keep LA-ICP-MS samples as close to the middle of the shell

555

section as possible, non-linear growth patterns could have caused variations in the actual location

556

of LA-ICP-MS samples hence affecting the element ratios [101], since the sample spot alignment

557

method used in this study [69] could not correct for measurement bias caused by variability in

558

CaCO3matrix. Furthermore, the sample alignment method assumed two-dimensional sampling

559

ignoring any effects of LA-ICP-MS sample volume. Consequently, the curvature of growth

560

lines deeper in the sample could have increased imprecision of element ratios through three-

561

dimensional time averaging. Despite all these uncertainties, our dataset is extensive and clearly

562

indicates that all of the studied elemental ratios were affected by several factors to the extent that

563

no element ratio in this study could be used as an absolute straightforward proxy of temperature,

564

salinity, fluorescence or shell growth rate.

565

5 Conclusions

566

We conclude that Ba/Ca, Li/Ca and Mg/Ca have a potential as environmental proxies inS.

567

groenlandicusandC. ciliatumshells: Incorporation of Ba/Ca might be connected with seasonal

568

dissolved or particular Ba dynamics in ambient water, and incorporation of Li/Ca and Mg/Ca

569

are likely connected with both CaCO3crystal growth rate and seawater temperature. Despite

570

this, all studied element ratios were likely affected by multiple internal and external factors

571

complicating the interpretation of element ratios. Our study was further affected by method-

572

ological constraints, such as time-averaging error, experimental artifacts, and uncertainties in

573

sub-annual growth models leading to partly inconclusive results for Sr/Ca and Mo/Ca. Despite

574

this our results are an important contribution to high-latitude bivalve shell geochemisty high-

575

lighting that none of the studied elemental ratios can be used as all-encompassing proxies of

576

seawater temperature, salinity, paleoproductivity, or shell growth rate. This, however, does not

577

preclude the use of element-to-calcium ratios as environmental proxies, but merely indicates that

578

seasonal dynamics of elements in seawater and seasonal variations in bivalve metabolism must

579

be understood better to link the elemental ratios in bivalve mollusk shells with environmental

580

processes.

581

(15)

Acknowledgments

582

We acknowledge the use of the NSF-supported WHOI ICP-MS facility and thank Scot Birdwhis-

583

tell for his excellent assistance. We are grateful to Bates Imaging Center and William Ash for

584

help with thick-section photographs. Further, we want to thank the Stack Exchange community

585

for help with the graphical presentation and data-analysis, and the R community for maintaining

586

open source statistics tools used in this study. This research was financed through the UiT The

587

Arctic University of Norway Utenlandstipend (MV), the EU 7th Framework Program project

588

Arctic Tipping Points (contract number FP7-ENV-2009-226248; http://www.eu-atp.org; PER),

589

the Research Council of Norway project Havet og Kysten (184719/S40; PER), the Norwegian

590

Polar Institute (HH, MV) and Akvaplan-niva (PER, MV, WGA, MLC).

591

References

592

1. Vaughan, D.; Comiso, J.; Allison, I.; Carrasco, J.; Kaser, G.; Kwok, R.; Mote, P.;

593

Murray, T.; Paul, F.; Ren, J.; Rignot, E.; Solomina, O.; Steffen, K.; Zhang, T. Ob-

594

servations: Cryosphere. InClimate Change 2013: The Physical Science Basis Contri-

595

bution of Working Group I to the Fifth Assessment Report of the Intergovernmental

596

Panel on Climate Change; Stocker, T.; Qin, D.; Plattner, G.K.; Tignor, M.; Allen, S.;

597

Boschung, J.; Nauels, A.; Xia, Y.; Bex, V.; Midgley, P., Eds.; Cambridge University

598

Press: Cambridge, UK and New York, NY, USA, 2013; chapter 4, pp. 317–382.

599

2. AMAP. Arctic climate issues 2011: Changes in Arctic snow, water, ice and permafrost.

600

SWIPA 2011 Overview report. Technical report, Arctic Monitoring and Assessment

601

Programme (AMAP), Oslo, Norway, 2012.

602

3. Grebmeier, J.M. Shifting patterns of life in the Pacific Arctic and sub-Arctic seas.Ann

603

Rev Mar Sci2012,4, 63–78.

604

4. Masson-Delmotte, V.; Schulz, M.; Abe-Ouchi, A.; Beer, J.; Ganopolski, A.; Rouco,

605

J.G.; Jansen, E.; Lambeck, K.; Luterbacher, J.; Naish, T.; Osborn, T.; Otto-Bliesner,

606

B.; Quinn, T.; Ramesh, R.; Rojas, M.; Shao, X.; Timmermann, A. Information from

607

paleoclimate archives. InClimate Change 2013: The Physical Science Basis Contri-

608

bution of Working Group I to the Fifth Assessment Report of the Intergovernmental

609

Panel on Climate Change; Stocker, T.; Qin, D.; Plattner, G.K.; Tignor, M.; Allen, S.;

610

Boschung, J.; Nauels, A.; Xia, Y.; Bex, V.; Midgley, P., Eds.; Cambridge University

611

Press: Cambridge, UK and New York, NY, USA, 2013; chapter 5, pp. 383–464.

612

5. Pearson, P.N.; Palmer, M.R. Atmospheric carbon dioxide concentrations over the past

613

60 million years. Nature2000,406, 695–699.

614

6. Zachos, J.; Pagani, M.; Sloan, L.; Thomas, E.; Billups, K. Trends, rhythms, and

615

aberrations in global climate 65 Ma to present.Science2001,292, 686–93.

616

7. Dame, R.F. Introduction. In Ecology of Marine Bivalves, An Ecosystem Approach;

617

CRC Press: Boca Raton, FL, USA, 2012; pp. 1–19.

618

Referanser

RELATERTE DOKUMENTER

In the following we consider three types of ratios of developments in the financial features of companies on the Oslo Stock Exchange: gearing ratios, interest cover- age ratios

In the following sections we will discuss the variational multiscale method as a turbulence modelling tool, and describe the implementation of the method in a spectral element

As natural variation of less than 3% in the geo-element ratios of 86 Sr: 87 Sr has been used to separate natal habitats in some fish species with up to 80% correct assignment

The ratio groupings were compared to recent reconstructions of seawater Mg:Ca and Sr:Ca ratios (3, 7) from younger Earth history suggesting that the increase in ratios during

A study on the remineralization ratios in the North Atlantic Ocean showed higher than Redfield C:nutrient ratios in the remineralized material in the deeper waters, and thereby a

The process for building the matrix of coefficients is normally based on traversing the given mesh (element by element), in order to identify the neighboring elements, which need

In order to attain a deeper understanding of how the Finite Element Method can be implemented in a parametric design environment, some Finite Element Analysis software packages

Effect of bone material properties on the initial stability of a cementless hip stem: a finite element study.. A CT-based high-order finite element analysis of the human